Skip to main content

Full text of "Analytical Rescaling of Polymer Dynamics from Mesoscale Simulations"

See other formats


Analytical Rescaling of Polymer Dynamics from Mesoscale Simulations 



I. Y. Lyubimov, J. McCarty, A. Clark, and M. G. Guenza* 
Department of Chemistry and Institute of Theoretical Science, University of Oregon, Eugene, OR 97403, USA 

We present a theoretical approach to scale the artificially fast dynamics of simulated coarse-grained 
polymer liquids down to its realistic value. As coarse-graining affects entropy and dissipation, 
two factors enter the rescaling: inclusion of intramolecular vibrational degrees of freedom, and 
rescaling of the friction coefficient. Because our approach is analytical, it is general and transferable. 
Translational and rotational diffusion of unentangled and entangled polyethylene melts, predicted 
from mesoscale simulations of coarse-grained polymer melts using our rescaling procedure, are in 
quantitative agreement with united atom simulations and with experiments. 



I. INTRODUCTION 

The development of a systematic approach to bridge 
time scales between different hierarchical levels of de- 
scription is an important goal in many areas of the 
physics of complex systems. pQ Furthermore, understand- 
ing the dynamics of polymeric liquids is relevant for many 
technological applications. Polymeric materials are pro- 
cessed in their liquid state, and the custom tailoring of 
new materials requires detailed predictions of their me- 
chanical and dynamical properties to be made based on 
their chemical structure. To this end, molecular dy- 
namics (MD) simulations shed light on the properties 
of complex systems; however, MD is limited by the pre- 
cision of the calculations, which degrades with the num- 
ber of computer iterations. [5] Quantitative predictions of 
the dynamics of unentangled polymer liquids are pos- 
sible through simulation runs that are computationally 
demanding because the polymer diffusion coefficient, D, 
scales with the degree of polymerization, N, as D oc TV -1 . 
Even more demanding are simulations of liquids of long, 
entangled, polymeric chains, where the diffusion coeffi- 
cient scales as D oc N~ 2 . For entangled polymer liquids 
it is therefore difficult to obtain well-equilibrated sam- 
ples or to simulate the system for several relaxation cy- 
cles, which would improve the precision of the calculated 
time-correlation functions. 

To reach the long time and length scales of interest, 
special strategies need to be employed. For example, 
some gain in computational time has been achieved by 
speeding up the equilibration process through an end- 
bridging Monte Carlo algorithm. [3] Moreover, if only 
qualitative, and not quantitative, predictions are sought, 
it is possible to adopt simplified intra- and intermolcc- 
ular potentials, which can speed up each simulation 
step,[H [5J or purely phenomenological coarse-grained 
descriptions, which reproduce the expected dynamical 
scaling behavior. [51 [5] However, quantitative computa- 
tional methods that precisely relate chemical structure 
to the long-time properties of the polymeric liquid are 



'Author to whom correspondence should be addressed. Electronic 
mail: mguenza@uoregon.edu 



still lacking. 

There is a need for computationally efficient, predictive 
methods to simulate polymer liquids and complex fluids 
in the long-time regime. The strategy we are pursuing 
here is to develop coarse-graining methods and dynami- 
cal rescaling procedures which start from first principles 
theory. [5] The goal is to obtain quantitative predictions 
of real polymer dynamics directly from properly rescaled 
fast mesoscale (MS) simulations of coarse-grained liquids. 

In a coarse-grained description, the system is specified 
by a set of relevant mesoscopic variables while smaller 
length scale variables are omitted. [9] This is done at the 
expense of entropy and dissipation (friction), which are 
underestimated by coarse-graining. [7] Because of the re- 
duced molecular degrees of freedom and the simplified 
energy landscape, MS-MD simulations require less com- 
putational time than atomistic simulations of the same 
systems. [HI H0HI1] However, because the effective energy 
landscape of the coarse-grained representation is artifi- 
cially smooth, MS-MD simulations predict accelerated 
dynamics, which need to be rescaled to produce realistic 
values. 

The enhanced diffusion in coarse-grained systems 
arises from the "soft" nature of the intermolecular poten- 
tial. This is advantageous in enabling larger time steps 
to be used in integrating the equations of motion, and an 
efficient sampling of the energy landscape, which leads to 
good statistical averages of the structural properties on 
the large scale. [13] However, the measured dynamics of 
coarse-grained systems is too fast, and as of yet, there is 
no clear procedure to quantitatively re-scale these accel- 
erated dynamics. 

In an effort to develop quantitative rescaling meth- 
ods, it is custom to build a numerical "calibration curve" 
obtained from direct comparison of MS-MD time cor- 
relation functions with atomistic simulations. The "cali- 
bration curves" are parametric, normal mode dependent, 
and specific to the system against which they are opti- 
mized, as well as to its thermodynamic conditions. Build- 
ing these parametric curves in part defeats the purpose of 
the coarse-graining procedure as it requires running many 
atomistic simulations to optimize the fitting parameters. 
If the level of coarse-graining is low, i.e. if the average is 
performed over a small number of atoms, the correction 
to the dynamics is only perturbative and the paramet- 



2 



ric rescaling works well.[T3] This explains the success of 
united-atom (UA) simulations. (3J [15] However, because 
the gain in computational time increases with the level 
of coarse-graining, simulations of slightly coarse-grained 
systems, e.g. UA-MD, afford a limited gain in time and 
length scales. 

The most gain in computational efficiency is achieved 
through large-scale coarse-graining, as the one adopted in 
this paper. This is most useful when studying bulk phys- 
ical quantities, e.g. viscosity, or systems with large-scale 
fluctuations, e.g. approaching spinodal decomposition. 
[16) Once simulations of heavily coarse-grained systems 
are combined with local scale simulations in a multiscale 
procedure, they provide the complete description of the 
system at all lengthscales of interest. [17] 

This paper presents a derivation of a first-principles 
approach to rescale the dynamics from MS-MD simu- 
lations of a coarse-grained polymer melt to the values 
of an atomistic description. The rescaling is analytical, 
general and transferible. The favorable comparison with 
atomistic simulations and experimental data in different 
thermodynamic conditions supports the validity of the 
proposed procedure. The theoretical basis of the rescal- 
ing rests on the fact that the accelerated dynamics is a 
result of the missing dissipation due to the eliminated 
degrees of freedom. 

The paper is organized as following. In Section 2 
we briefly review our coarse-grained model, while the 
coarse-grained potential and mesoscale simulations are 
described in Section 3. In the following section we present 
our atomistic model and the calculation of the energy 
due to the internal degrees of freedom. Rescaling of 
the friction coefficient is discussed in Section 5, followed 
by a comparison of the diffusion coefficients predicted 
from MS-MD simulations after rescaling, with united- 
atom simulation and experimental data. A brief discus- 
sion concludes the paper. 



The form factors entering Eq.Q are approximated by 

uj cm (k) = iVe'T and uj mm {k) = N/(l + k 2 R 2 g /2), 
which is the Pade approximant of the Debye function 
uj mm {k) = 2N(e- k2R l + k 2 R 2 g - l)/(fc 4 i?4). For the 
monomer-monomer total intermolecular correlation func- 
tion h mm (k), we use the thread-limit polymer reference 
interaction site model description, [T!5] in which h mm (k) — 
W(l + £p fc2 )(! + k2 &- Here ; £p is tlie length scale of 
density fluctuations defined as t;^ 1 = + ^ p _1 , with 
£ c = Rg/y/2 the length scale of the correlation hole, and 
£' p = R g /(2irp* s ) with p* = p s R 3 g being the reduced molec- 
ular number density. The number density of soft colloidal 
particles p s = p/N with p being the monomer density 
and h = (e p /e - 1)1 Ps- 

The structure of the coarse-grained polymer liquid is 
described by the total distribution function, [8] approxi- 
mated for polymer chains with N > 30 as 

which results from the analytical Fourier transform of 
h(k). The structure of the liquid on length scales of the 
order of the polymer radius-of-gyration and larger, is well 
described by Eq.Q, which is in quantitative agreement 
with both atomistic and coarse-grained simulations.^ 
The coarse-grained description of Eq.|2]) is thermody- 
namically consistent with the atomistic representation, 
e.g. of the liquid compressibility. Because the total 
correlation function of the coarse-grained representation 
is analytical, i.e. it depends explicitly on density and 
molecular parameters, it is also general and state-point 
transferable. 



II. COARSE-GRAIN MODEL 

Our coarse-grained representation models polymers as 
interacting soft-colloidal particles with repulsive interac- 
tion of the order of the size of the macromolecule, de- 
fined by the radius-of-gyration, i? ff .[B] The total distribu- 
tion function is derived from the solution of a generalized 
Ornstein-Zernikc equation, treating coarse-grained sites 
as auxiliary sites, and atomic sites as real sites|18j. In 
reciprocal space, the total distribution function is 



h(k) = [w cm (fc)/cj r 



l (k)fh r - 



\k) 



(1) 



where the superscript "mm" identifies the monomer- 
monomer distribution, while "cm" indicates the distri- 
bution of monomers with respect to the center-of-mass. 
Eq.(T]) is solved by assuming a Gaussian description 
of the intramolecular site distribution, which is an ac- 
cepted approximation for polymers in a liquid state. 



III. COARSE-GRAINED POTENTIAL AND 
MESOSCALE SIMULATIONS 

Each soft-colloidal particle interacts with other colloids 
through an effective potential of the range of the overall 
polymer dimension, R g . While hard-sphere systems are 
best described by a Percus-Yevick closure, the Hyper- 
Netted Chain (HNC) closure works best for systems with 
soft potentials, [20] including the mesoscopically coarse- 
grained polymer melts investigated here, [HI [21] aud poly- 
mer coils in dilute or semidilute solutions. [H] The HNC 
potential between a pair of coarse-grained units is de- 
rived by applying the closure j3v{r) — h(r) — ln[h(r) + 
1] — c(r), where the direct correlation function is de- 
fined by the Ornstein-Zernike relation in reciprocal space 
c(k) = h(k)/(l +M(fc)).[23] 

The potential between two spheres is calculated nu- 
merically, after Fourier transform, from the analytical 
expression, Eq.Q, by adopting the Debye form of the 



3 



monomer intramolecular distribution, u> mrn (k). The De- 
bye approximation has been shown to better represent 
simulation data than its Pade approximant. 2lj Tabu- 
lated HNC potentials are input to the mesoscale simula- 
tions. 

MS-MD simulations of polymer liquids are performed 
in the microcanonical ensemble, where each molecule is 
represented as an interacting soft-colloidal particle. In 
the initialization step, all particles are placed on a lattice 
with periodic boundary conditions. Each site is given an 
initial velocity and subsequently, the system is evolved 
using a velocity Verlet integrator. Equilibrium is induced 
in the ensemble by rescaling the velocity at regular in- 
tervals until the desired average temperature is reached. 
At this stage, velocity rescaling is discontinued and tra- 
jectories are collected over a traversal of ~ 8R g , while 
the temperature is monitored to assure that it fluctuates 
around the desired equilibrium value. Because of the 
form of Eq. ^ , the simulation uses reduced quantities of 
distance, R g = 1, mass, m = 1, and energy, ksT = 1. 

A typical MS-MD simulation for our model is per- 
formed on a single-CPU workstation, and consists of 
~ 3000 polymers, evolving for a duration of ~ 4 hours. 
The MS-MD simulation provides identical structural in- 
formation to that of the analogous atomistic or united- 
atom (UA) simulations on length scales equal and larger 
than the polymer R g . However, the MS-MD requires a 
much smaller computational power than the atomistic 
and UA-MD simulation, which is typically performed on 
a liquid of ~ 500 polymers, on a parallel supercomputer. 
The convenient requirements in computational power of 
the MS-MD simulation allows one to increase consider- 
ably the number of particles and the simulation box size 
without dramatically affecting the computational time, 
thus improving the precision on the large-scale data col- 
lected. 



IV. MAPPING OF THE ATOMISTIC 

DESCRIPTION ONTO A 
FREELY- ROTATING-CHAIN MODEL 

The correction in Hclmhotz free energy, which accounts 
for the discarded internal degrees of freedom, is calcu- 
lated starting from the atomistic representation of the liq- 
uid, where each chain is described as a collection of beads 
connected by springs defined by an effective intramolecu- 
lar quadratic potential U (r) = 3ksT/ (2l 2 ) X)fj=i ^-i,j r i ' 
Tj. Here A is the connectivity matrix, which represents 
the structure and local flexibility of the polymer, the 
position of unit i in a chain of N beads, and 1^ = r j+i — 
the bond vector connecting two adjacent beads. [24] For 
polyethylene melts, each polymer is represented as a 
freely-rotating-chain (FRC), finite in size, with semiflcx- 
ibility parameter g = ((1, • = 0.785 and 

fixed bond length I — 1.54A. This bead-and-spring model 
of the FRC has been shown to represent correctly the dy- 
namics of polyethylene melts as measured in united-atom 



simulations [23] and in experiments. [26] A FRC model 
was also successfully adopted to model the dynamics of 
polymer melts with different molecular architectures. [2"T] 
and even proteins, [M] [25] once the proper semiflexibility 
parameters are selected. 

To map the MS-MD simulation onto a real system, the 
reduced unit of time needs to be properly rescaled. Be- 
cause energy is dissipated in internal degrees of freedom 
in the atomistic representation, the contribution due to 
the internal vibrational modes is included in the coarse- 
grained representation by rescaling the time in the MS- 
MD simulation, t, by the amount of internal free energy 
dispersed in vibrational modes in the atomistic descrip- 
tion as t — tR g y/3mN/(2kBT), with the particle mass, 
m, and size R g . By rescaling the internal free energy, we 
accounts for the change in entropy in the coarse-grained 
description. In the next section we derive the friction 
rescaling. 

V. RESCALING OF THE FRICTION 
COEFFICIENT 

To account for the change in dissipation caused by 
coarse-graining, we start from the diffusion coefficient 
measured in the MS-MD simulation, D^ IS , and we de- 
rive the center-of-mass (cm) diffusion coefficient of the 
polymer, D cmi through the rescaling of the friction as 

D cm = D? IS ( s /(N( m ) . (3) 

The correction factor is calculated from the ratio of the 
cm friction in the soft colloid representation £ s and the 
cm friction in the atomistic representation N£ m , with 
Cm the monomer friction. Each friction coefficient is 
evaluated by solving the memory function in the corre- 
sponding Generalized Langevin Equation. These mem- 
ory function definitions result from the straightforward 
application of Mori-Zwanzig projection operators to the 
Liouvillc equation, where either the monomer (atomistic 
description) or the center-of-mass (soft colloid descrip- 
tion) of the polymer are assumed to be the "relevant 
slow variables." [21 [M] 

In the soft colloid representation the friction coefficient 
is defined as 

Q={p s r dt I dv f dv'g(r)g(r')F(r)F(r')v-v' 

6 Jo J J ^ 

x J dRS(R;t)S(\r-r' + R|;t), 

where f3 = 1/kgT, g(r) = h(r) + 1 is the radial dis- 
tribution function, F(r) = (din g(r) / 'dr) is the total 
force exerted by the surrounding fluid on the colloid, and 
S(k) = 1 + p s h(k) is the structure factor of the fluid sur- 
rounding the colloid. The unit vectors f and f ' define the 
direction of exerted forces. 



4 



In Eq.Q, the projected dynamics has been substituted 
with the real (unprojected) dynamics, which is a valid 
approximation when the Langevin equation is expressed 
as a function of slow variables for the diffusive regime. [53] 
In the long-time regime, the relaxation of the liquid is 
dominated by the polymer ccnter-of-mass diffusion, D, 
here represented by the cm of the colloidal particle. In 
Fourier space the dynamic structure factor of the liquid 
reads S(k; t) ~ S(k)exp(—k 2 Dt). Evaluating the integral 
with use of Eq.Q gives 



D/3( s KAyfrp,R g § ( 1 + f- 



1183 
~507 



p s h 



507 
512 

679^ 2 ,o 

piK 

1024 Hs 



(5) 



Eq.© expresses the friction coefficient of a soft col- 
loidal particle. 

In the atomistic description, the monomer friction co- 
efficient is defined by the memory function as 



E f dT \p J dr J dr'g(r)g(r')F(r)F(r') 

j,i=l 

xf J dRSij{R; t)S(\r - r' + R|; t) , 

(6) 

where the dynamic structure factor of the surrounding 
liquid is approximated as S(k;t) S{k)exp{—k 2 Dt) = 
[ui(k) + ph(k)]exp(—k 2 Dt), with uj(k) being the in- 
tramolecular static structure factor. This expression for 
S(k;t) assumes that in the long time regime the relax- 
ation of the liquid is dominated by the polymer center- 
of-mass diffusion, consistently with the soft-colloid rep- 
resentation. 

To evaluate Eq.(|6|, we approximate the potential as 
an effective hard-core potential with a diameter d to be 
defined. [2"5] In hard-core fluids, g(r)F(r) = g(d)j3~ 1 8(r — 
d). Working in reciprocal space, the integrals in Eq. ([6| 
can be performed analytically to give an expression for 
the dynamical quantity D(3C, m depending on two length 
scales: R g , and d. The result is lengthy, and it is not 
reported here. [53] The values of R g used are the ones 
reported in Tables [I] and [rJ] 

The monomer hard-core diameter, d = 2.1 A, is identi- 
cal for all the samples, and is obtained by reproducing the 
scaling with N of an unentangled melt, Z?/3£ m = 1/N, for 
the PE 44 sample. This sample is chosen because its de- 
gree of polymerization is smaller than the entanglement 
one, N e — 130, and it is large enough to ensure Gaussian 
chain statistics. Once d is defined, it is not changed for 
any other system considered, either unentangled or en- 
tangled. This is the only parameter that has to be fixed 
in our approach. Theoretically predicted values for un- 
entangled systems recover the correct scaling behavior as 



ND(3( m w 1. For entangled systems, we solve Eq.([6]) by 
including a one-loop perturbation of the diffusion coeffi- 
cient. For these entangled systems ND(3( m cx A -1 , in 
agreement with the known scaling behavior. 



VI. COMPARISON OF PREDICTED 
DIFFUSIVE DYNAMICS WITH SIMULATIONS 
AND EXPERIMENTS 

To test our approach we compare the rescaled dynam- 
ics, predicted from mesoscale simulations, with experi- 
ments and UA-MD simulations. Each sample that we 
investigate is in well-defined thermodynamic conditions 
of density and temperature, and has a specific radius 
of gyration. Those quantities enter as an input to our 
mesoscale simulation, and also in the expressions for the 
rescaling of the energy and friction coefficient (see Tables 
|l]and|n|. 

Comparison with simulation data are limited here to 
United-Atom simulations, but our theory is general and 
comparison could be made with atomistic simulations 
as well. The UA-MD simulations reported in this pa- 
per cover a regime from unentangled, [15 to slightly en- 
tangled dynamics (two entanglements per chain). [3] We 
use as an input of our approach the radius of gyration, 
as measured in each simulation. These values are very 
close to the theoretical R g values calculated using a FRC 
model with semiflcxibility parameter g — 0.785. 

The experimental samples considered in this paper [26 
cover a region at the crossover from unentangled to en- 
tangled dynamics comparable to the one in UA-MD sim- 
ulations. However, the values of the radius-of-gyration 
for those samples are not known, because only the de- 
gree of polymerization is reported. For those samples we 
assume as input values of R g those calculated using a 
FRC approach. 

The predicted cm diffusion coefficient of a polymer 
chain, D cm is compared in Figure [T] against the data from 



TABLE I: MS-MD Parameters - UA-MD data 



Polymer 


N 


T [K] 


p [sites/A 3 ] 


R g [A] 


PE30 a 


30 


400 


0.0317 


7.97 


PE44 a 


44 


400 


0.0324 


10.50 


PE48 b 


48 


450 


0.0314 


10.54 


PE66 a 


GG 


448 


0.0329 


13.32 


PE78 b 


78 


450 


0.0321 


14.35 


PE96 a 


96 


448 


0.0328 


16.79 


PE142 b 


142 


450 


0.0327 


20.51 


PE174 b 


174 


450 


0.0328 


22.92 


PE224 b 


224 


450 


0.0329 


26.28 


PE270 b 


270 


450 


0.0330 


29.27 


PE320 b 


320 


450 


0.0330 


31.31 



° data from ref. [15]; b data from ref. [3] 



5 



TABLE II: MS-MD Parameters - Experimental Data [T = 
509K, p = 0.0315 [sites/A 3 ]. Data from Ref.|26]] 



Polymer TV i^ flC [A] 

PE36 36 10.07 

PE72 72 14.82 

PE106 106 18.20 

PE130 130 20.25 

PE143 143 21.27 

PE192 192 24.77 

PE242 242 27.88 



simulations, [51 [TS] and from experiments. jUj We also 
show, as a guide to the eye, lines with the scaling behavior 
of unentangled and entangled systems. The agreement 
between calculated and measured diffusion coefficients is 
good over a range of the degree-of-polymerization, cov- 
ering unentangled as well as entangled polymer melts. 
Because each simulation and experimental value is taken 
in slightly different thermodynamic conditions, the data 
points do not perfectly align along the lines of the scal- 
ing exponents in the figure. However we observe a good 
agreement between predicted theoretical values and mea- 
sured ones in simulations or experiments. 



i<r 

Q 

10° 


: (a) 


i - 

i i' 




: (b) 




io 2 






SO 

.d 

^10' 
Q 




t : 


10° 




1 I : 



io 2 io 3 

N 



FIG. 1: Plot of diffusion coefficients as a function of degree 
of polymerization, TV. Comparison between the theoretically 
predicted values (triangle), simulations (square) from [51 115] . 
and experiments (circle) from [26] and references therein. 
Also shown is the scaling for unentangled, TV" 1 (dot-dashed 
line), and entangled systems, TV -2 (dashed line) 

To test further the validity of our procedure, we calcu- 
late the decay of the rotational time-correlation function 
for the molecular end-to-end vector with input parame- 



ters for polyethylene and the rescaled monomer friction 
coefficient £ = fc^T '/ '(TV D cm ), and compare it against 
UA-MD simulations (see Figure [2]). Predicted and mea- 
sured decays are in excellent agreement for the unen- 
tangled samples, suggesting that the proposed procedure 
holds for different normal modes of motion. 




o 2 4 6 8 10 
t[ns] 



FIG. 2: Normalized rotational time decorrelation function for 
the end-to-end vector for the semiflexible chains with rescaled 
friction (solid lines) compared against simulations (symbols) 
for polyethylene melts of increasing length; TV = 30 (circles), 
TV = 66 (squares), TV = 96 (triangles) 



VII. CONCLUSIONS 

Mesocale simulations of coarse-grained systems are be- 
coming increasingly important, as they are computa- 
tionally efficient and allow for the study of systems on 
larger length and time scales than their atomistic coun- 
terparts. However, while structural properties on large 
length scales are well described by MS-MD simulations, 
the dynamics is unrealistically fast due to the simplified 
free energy landscape. In this paper we have presented 
an analytical, first-principles, approach to scale the dy- 
namics measured in mesoscale simulations down to re- 
alistic atomistic values. The rescaling procedure takes 
into account the averaged internal degrees of freedom and 
enhanced dissipation due to the coarse-grainining proce- 
dure. The agreement of predicted long-time dynamics 
with data from simulations and experiments is quanti- 
tative. Whereas previous efforts at dynamical rescaling 
have used numerical calibration curves, which are specific 
of the system under study, our approach is analytical and 
thus general and transferable: it is readily applicable to 
systems with different thermodynamic parameters and 
to polymer chains of increasing degree of polymerization 
crossing from the unentangled to the entangled regime. 
The development of a general scheme to rescale the dy- 
namics from MS-MD simulations promises to be useful 
in multiscale modeling techniques and fast equilibration 



6 



methods employed in computer simulations of complex 
fluids. 

In summary, the development of schemes to rescale the 
dynamics from MS-MD simulations will certainly be ben- 
eficial in the advancement of multiscale modeling tech- 
niques of complex fluids. Equilibrium and non equilib- 
rium simulations of polymer melts with different architec- 
tures should be natural implementations of the approach 
for future work. It should also be possible to extend this 
model to rescale systems represented at an intermedi- 
ate level of coarse-graining as collections of soft colloidal 
beads, so that the internal dynamics of entangled chains 



can be simulated. 



ACKNOWLEDGNEMTS 

We acknowledge support from the National Science 
Foundation. Simulation trajectories for the UA-MD sim- 
ulations were kindly provided by Gary G. Grest and Vla- 
sis G. Mavrantzas. We thank Glenn T. Evans for the 
careful reading of the manuscript and helpful suggestions. 



[1] M. Gell-Mann and J. B. Hartle, Phys. Rev. A 76, 022104 [16 
(2007). 

[2] D. Frenkel and B. Smith, Understanding Molecular Sim- [17 
ulation. From Algorithms to Applications (Academic 
Press, London, 2002). [18; 

[3] A. Uhlherr, M. Doxastakis, V. G. Mavrantzas, D. N. 

Theodorou, S. J. Leak, N. E. Adam and P. E. Nyberg, [19 
Europhys. Lett. 57, 506 (2002). 

[4] M. Kroger and S. Hess, Phys. Rev. Lett. 85, 1128 (2000). [20 

[5] S. K. Sukumaran and A. E. Likhtman, Macromolecules 
42, 4300 (2009). 

[6] J. T. Padding, and W. J. Briels, J. Chem. Phys. 117, [21 
925 (2002). 

[7] H. C. Ottinger Beyond Equilibrium Thermodynamics [22 

(Wiley, Hoboken, N.J.2005). 
[8] G. Yatsenko, E. J. Sambriski, M. A. Nemirovskaya and [23 

M. Guenza, Phys. Rev. Lett. 93, 257803 (2004). 
[9] M. Karttunen, I. Vattulainen, and A. Lukkarinen (eds.), [24 
Novel Methods in Soft Matter Simulations; Lect. Notes 
Phys. 640 (Spinger-Verlag Berlin Heidelberg 2004). [25 
[10] T. A. Knotts IV, N. Rathore, D. C. Schwartz and J. J. [26 

de Pablo, J. Chem. Phys. 126, 084901 (2007). 
[11] M. L. Klein and W. Shinoda, Science 321, 798 (2008). 
[12] G. Milano and F. Muller-Plathe, J. Phys. Chem. B 109, [27; 
18609 (2005) 

[13] S. O. Nielsen, C. F. Lopez, G. Srinivas and M. L. Klein [28 

J. Phys.: Condens. Matter 16, R481 (2004). 
[14] V. A. Harmandaris and K. Kremer, Macromolecules 42, 

791 (2009). [29; 
[15] M. Mondello and G. S. Grest, J. Chem. Phys. 106, 9327 

(1997). 



J. McCarty, I. Y. Lyubimov and M. G. Guenza, Macro- 
molecules 43, 3964 (2010). 

J. McCarty, I. Y. Lyubimov and M. G. Guenza, J. Phys. 
Chem. B 113, 11876 (2009). 

V. Krakoviack, J. -P. Hansen and A. A. Louis, Europhys. 
Lett. 58, 53 (2002). 

K. S. Schweizer and J. G. Curro, Adv. Chem. Phys. 98, 
1 (1997). 

McQuarric, D. A. Statistical Mechanics; University Sci- 
ence Books: Sausalito, C. A., 2000 (see discussion in Sec- 
tion 13-9). 

E. J. Sambriski, G. Yatsenko, M. A. Nemiroskaya and M. 
G. Guenza, J. Chem. Phys. 125, 234902 (2006). 
A. A. Louis, P. G. Bolhuis, J. P. Hansen and E. J. Meijer, 
Phys. Rev. Lett. 85, 2522 (2000). 

J. -P. Hansen and I. R. McDonald, Theory of Simple Liq- 
uids (Academic Press, London, 1991). 
M. G. Guenza, J. Phys.: Condens. Matter 20, 033101 
(2008), and references therein. 
M. Guenza Phys. Rev. Lett. 88, 25901 (2002). 
M. Zamponi, A. Wischnewski, M. Monkenbusch, L. Will- 
ner, D. Richter, P. Falus, B. Farago and M. G. Guenza J. 
Phys. Chem. 112, 16220 (2008), and references therein. 
E. J. Sambriski, G. Yatsenko, M. A. Nemiroskaya and M. 
G. Guenza J. Phys.: Cond. Matt. 19, 205115 (2007). 

E. Caballero-Manrique, J. K. Brey, W. A. Deutschman, 

F. W. Dahlquist and M. G. Guenza Biophys. J. 93, 4128 
(2007). 

The mathematical notebook with the solution is available 
upon request.